You are on page 1of 13

Astronomy & Astrophysics manuscript no.

8322main February 2, 2008


(DOI: will be inserted by hand later)

Effects of photophoresis on the evolution


of transitional circumstellar disks
Fabian Herrmann1 and Alexander V. Krivov1
arXiv:0711.2595v1 [astro-ph] 16 Nov 2007

Astrophysikalisches Institut, Friedrich-Schiller-Universität Jena, Schillergäßchen 2–3, 07745 Jena, Germany

Received July 20, 2007, accepted 1 October 2007

Abstract. Although known for almost a century, the photophoretic force has only recently been considered in astrophysical
context for the first time. In our work, we have examined the effect of photophoresis, acting together with stellar gravity,
radiation pressure, and gas drag, on the evolution of solids in transitional circumstellar disks. We have applied our calculations to
four different systems: the disks of HR 4796A and HD 141569A, which are several Myr-old AB-type stars, and two hypothetical
systems that correspond to the solar nebula after disk dispersal has progressed sufficiently for the disk to become optically thin.
Our results suggest that solid objects migrate inward or outward, until they reach a certain size-dependent stability distance
from the star. The larger the bodies, the closer to the star they tend to accumulate. Photophoresis increases the stability radii,
moving objects to larger distances. What is more, photophoresis may cause formation of a belt of objects, but only in a certain
range of sizes and only around low-luminosity stars. The effects of photophoresis are noticeable in the size range from several
micrometers to several centimeters (for older transitional disks) or even several meters (for younger, more gaseous, ones). We
argue that due to gas damping, rotation does not substantially inhibit photophoresis.

Key words. planetary systems: formation – planetary systems: protoplanetary disks – circumstellar matter – celestial mechanics
– stars: individual: HR 4796A, HD 141569A.

1. Introduction 2005; Jayawardhana et al. 2006; Balog et al. 2007; Currie et al.
2007).
The standard (core accretion) scenario of planet formation Transition from gas- and dust-rich, optically-thick pro-
implies continuous growth of solids in a protoplanetary disk toplanetary disks to nearly gas-free, optically-thin debris
around a young star (see, e.g. Safronov 1972; Wetherill disks is currently in the focus of interest of both observa-
1980; Shu et al. 1987; Lissauer 1993; Strom & Edwards tional and theoretical effort (Wieneke & Clayton 1983; Morfill
1993; Weidenschilling & Cuzzi 1993; Papaloizou et al. 1999; 1983, 1988; Weidenschilling & Cuzzi 1993; Strom et al.
Blum & Wurm 2000; Wurm & Blum 2006; Henning et al. 1993; Zuckerman & Becklin 1993; Simon & Prato 1995;
2006; Meyer et al. 2007). Micron-sized dust grows step by Meyer & Beckwith 2000; Ardila et al. 2005; Calvet et al. 2005;
step to larger bodies, until asteroid-sized planetesimals are Hueso & Guillot 2005; Augereau 2006; Bouwman et al. 2006;
reached. At sizes well below 1 km, gravitational interac- Eisner et al. 2006). Examples of transitional objects are TW
tions between the objects play a minor role and the pro- Hya, HR 4796A and HD 141569A, while the well-known sys-
cess is largely determined by interactions of solids with the tem β Pic already enters into the realm of almost gas-free de-
ambient gas in the disk. Gas causes sedimentation, mix- bris disks (Thébault & Augereau 2005). Resolved transitional
ing, radial drift (Weidenschilling 1977; Kley et al. 1993; disks exhibit radial structure in the form of rings and gaps, of-
Takeuchi & Artymowicz 2001; Brauer et al. 2007) and other ten alternating, and there is a lot of debate whether this struc-
effects that all dictate the spatial, size, and velocity distribu- ture is caused by gravity of hidden planets (Augereau et al.
tions of grains and therefore the conditions for their growth. 1999b; Wyatt et al. 1999; Telesco et al. 2000) or by interac-
As the disk evolves, both gas and dust are gradually removed tion of solids with the ambient gas component (Klahr & Lin
from the systems. In this process, the gas-to-dust ratio re- 2000; Takeuchi & Artymowicz 2001; Besla & Wu 2007). The
duces from ∼ 100 in the initial protoplanetary disk stage, at first possibility is supported by the fact that the amount of gas
the ages on the order of 1 Myr, to vanishing values of ≪ 1 remaining in these systems is probably no longer sufficient to
at the debris disk stage after ∼ 10 Myr (Lawson et al. 2004; form Jupiter’s gas envelope (Chen & Kamp 2004) and so, the
Haisch et al. 2005; Hollenbach et al. 2005; Takeuchi et al. planet formation process must already be finished. The sec-
ond hypothesis is substantiated by simulations based on the
Send offprint requests to: Fabian Herrmann, e-mail: observational estimates of the gas contents. These show that
fabian@astro.uni-jena.de gas drag, acting together with other forces – stellar gravity and
2 Herrmann & Krivov: Photophoresis in transitional disks

radiation pressure — could result in segregation of different- eralize to the presence of photophoresis. To solve the equa-
sized solids and radial fractionation of dust. tion of motion, we employ a modification of the semianalytic
In this paper, we study the motion of different-sized solids scheme used by Weidenschilling (1977). The results are ap-
under the combined action of stellar gravity, radiation pressure, plied to circumstellar disks with different gas density around
and gas drag, to which we add photophoresis, an additional stars of different luminosity.
radial force that acts on particles in a gas disk which are ex- In Sect. 2, we discuss the astrophysical status and parame-
posed to the radiation field of the central star. Photophoresis, ters of the systems chosen for this study, as well as the formu-
one of the lesser known forces of physics, has been first de- las used for gas density and temperature in the disks. Section 3
scribed by Ehrenhaft (1917). The force is caused by the fol- provides formulas for different perturbing forces, including the
lowing process. If an object is embedded in a thin gas and photophoretic one. Section 4 deals with the equation of motion.
is exposed to an anisotropic radiation field (i.e. of a star), a In Sect. 5, we analyze the properties of the particles’ radial mo-
gas molecule that becomes accommodated by the object’s sur- tion and derive a formula for the radii of stable orbits. We also
face and rejected again will depart from the illuminated – and estimate the size ranges in which photophoresis should be taken
therefore warmer – side on average with a greater velocity into account. In Sect. 6, we examine the problem of particle ro-
then from the dark, colder one. Net momentum is transferred tation in order to check whether it might nullify the effect of
to the object, accelerating it away from the light source (note photophoresis. In Sect. 7, the results are summarized and ideas
that for very small particles, the force can also be attractive – for future research are presented.
see Tehranian et al. 2001). Although known for almost a hun-
dred years and successfully used in technical applications, such
2. The systems
as the construction of optical traps (Steinbach et al. 2004), it
has only recently been analyzed in astrophysical context. The Following Takeuchi & Artymowicz (2001), we choose two
first attempt to examine its influence on planetary formation transitional disks of particular interest for our study: HR 4796A
was made by Krauss & Wurm (2005). In their subsequent pa- and HD 141569A. We use these two systems in our work as
per (Wurm & Krauss 2006), they investigated how photophore- model environments for the forces we wish to explore, as their
sis might affect the formation of chondrules and the survival properties are quite well known and reasonably accurate mod-
of Calcium/Aluminum-rich inclusions. Finally, Krauss et al. els exist for the physical conditions (gas density and tempera-
(2006) analyzed the effects imposed by photophoresis on the ture) in their disks.
inner rim of a dusty protoplanetary disk. They assumed the disk The A0-type star HR 4796A, located at a distance
to stay optically thick at all times, thus allowing only its in- of 67.1+3.5
−3.4 pc from the sun (Hipparcos data), is a mem-
ner edge to be influenced by photophoresis. Using an exponen- ber of the TW-Hydrae-Association (TWA in the follow-
tial law to describe particle growth and the α 1+1D turbulent ing, see Kastner et al. 2001). According to current knowl-
model (Shakura & Sunyaev 1973) used in Alibert et al. (2005) edge, this is the young star association closest to our sun
and Papaloizou & Terquem (1999) for the evolution of the gas (Zuckerman & Song 2004). The age of TWA can be estimated
disk, Krauss and Wurm computed the motion of the inner edge. to be approximately 10 Myr. Since it has used up its molecular
This is identical to the motion of the smallest particles, since, cloud completely, the only gas left in in the association is bound
as will be shown subsequently, for solar luminosity systems, in the circumstellar disks. The stellar parameters of HR 4796A
the velocity of outward motion of an object increases with its are estimated to be M⋆ = 2.5 M⊙ , L⋆ = 21.0 L⊙ (Koerner et al.
size in a certain size range (compare Fig. 3). In essence, they 1998; Jayawardhana et al. 1998; Telesco et al. 2000). IRAS
showed that the dust disk’s edge first moves outward, pushed discovered that it emits 0.5 % of its entire radiative power in the
by photophoresis. After a few million years of time, it shrinks infrared part of the spectrum, which points to a high dust den-
towards the star again, because gas pressure declines as disk sity in its neighborhood – in fact, HR 4796A is the dust-richest
dispersal proceeds. star in the Bright Star Catalog. Stauffer et al. (1995) estimate
This paper focuses on transitional disks. In these systems, its age to be 8 ± 2 Myr, which makes it slightly younger than β
the conditions are quite favorable for photophoresis to be effi- Pic (12 Myr). The structure of the dust disk around HR 4796A
cient. Transitional disks are already optically thin, so that the is highly complicated. First resolved images in thermal infrared
particles are exposed to the strong radiation field. On the other were obtained by Koerner et al. (1998) and Jayawardhana et al.
hand, they are still sufficiently gas-rich. We would like to check (1998). Wahhaj et al. (2005) conducted elaborate studies of
whether, and to which extent, photophoresis may affect the ra- the disk’s structure, using data from MIRLIN at the Keck II
dial migration of solids. telescope as well as 350 µ m measurements from the Caltech
We use a single-body dynamics approach to investigate Submillimeter Observatory and Hubble Space Telescope scat-
the behavior of a two-dimensional swarm of particles, the tered light images. Their studies point to a disk composed of an
sizes of which are distributed in an interval ranging from inner, exozodiacal dust ring located at roughly r = 4 AU from
several micrometers to several meters. The size of an indi- the star and a wide, two-component outer dust belt consisting
vidual particle, though, is being kept constant. We develop of a broad ring of ∼ 7 µ m grains stretching from 45 to 125
a simple theory for the long-term behavior of a distribu- AU, and a narrower structure between 66 and 80 AU consist-
tion of solids experiencing gravity, radiation pressure, gas ing of ∼ 50 µ m grains. While the exozodiacal dust may be the
drag and photophoresis. Methodically, our study is similar to product of an asteroid-type belt, the outer belt can be explained
Takeuchi & Artymowicz (2001), whose computations we gen- most naturally by the assumption that the grains are emitted by
Herrmann & Krivov: Photophoresis in transitional disks 3

Table 1. Parameters of gas disk models.

System Example ρ0 [10−10 kg m−3 ] Density exponent q Luminosity L⋆ /L⊙ Mass M⋆ /M⊙
lGlL ‘older’ SN 3.12 -2.75 1.0 1.0
hGlL ‘younger’ SN 156 -2.75 1.0 1.0
lGhL HR 4796A∗ 3.12 -2.75 21.0 2.5
hGhL HD 141569A∗ 156 -2.25 22.4 2.3

∗ Sources: Takeuchi & Artymowicz (2001), their Table 1

an exo-Kuiper swarm of planetesimals. Since the smallest ones For gas density and temperature in all four model systems,
are diffused rapidly by radiation pressure, they form a broad we use standard power-law approximations (Hayashi et al.
belt, while the larger ones stay closer to their area of origin. 1985):
Also, asymmetric structures have been reported (Telesco et al.  1/4 
2000). They may be caused by secular perturbations of one or T L⋆ r −1/2
= 278 (1)
several planets or by the stellar companion HR 4796B. Most 1K L⊙ 1 AU
probably, the disk is perturbed by planetary bodies as well as and
by the B component. In order to define the mass and number of ρg  r q
planets, the orbit of HR 4796B would have to be known with = ρ 0 , (2)
1 kg m−3 1 AU
greater precision.
where ρ0 and q are constants that vary from one system to an-
other. To keep the treatment simple, we assume the exponent q
The second object, HD 141569A, has the Hipparcos to be constant throughout each disk and do not use a ‘break-off
distance of 99 ± 10 pc. Its parameters are estimated function’ to describe the density drop beyond a certain radius
as M⋆ = 2.3 M⊙ , L⋆ = 22.4 L⊙ (Jura et al. 1993, 1998; rout (Takeuchi & Artymowicz 2001). As will be shown subse-
van den Ancker et al. 1998) It is underluminous for its spectral quently, the outer part of the radial profile does not really mat-
type B9.5 Ve, which is a common occurrence for young A-type ter, because photophoresis affects only the inner parts of the
stars, found also at HR 4796A, β Pic and 49 Ceti (Jura et al. disk.
1998; Lowrance et al. 2000). Being, like HR 4796A, part of Altogether, our model contains three basic parameters: the
a multiple system with an M2 and an M4 component, the gas density ρ0 at r = 1 AU, the gas density exponent q which
lower-mass companions which have not yet reached the main controls the size of the gas disk, and the stellar luminosity L⋆ .
sequence can be used to find an estimate for the system’s age of Our four model systems essentially explore the photophoretic
5 ± 3 Myr (Weinberger et al. 2000). As in the HR 4796A case, effect in the ‘parameter rectangle’ luminosity – gas density:
the disk itself has a complex morphology (Weinberger et al.
1999; Augereau et al. 1999a; Fisher et al. 2000; Mouillet et al. – lLhG (low luminosity, high gas content, corresponding to
2001), consisting of two dust belts with radii of 200 and 325 the solar nebula at an earlier stage),
AU, the centers of which are shifted by 20 – 30 AU in the – lLlG (low luminosity, low gas content, solar nebula at a later
direction of the system’s semiminor axis. Between them a stage),
dust-free gap can be found at ∼ 250 AU. Also, inside of 150 – hLhG (high luminosity, high gas content, HD 141569A),
AU the disk’s luminosity declines to the level of background and
noise, which points to a strong depletion of dust in the system’s – hLlG (high luminosity, low gas content, HR 4796A).
inner region. The outer ring shows a tightly-wrapped spiral
The parameters of all four systems are listed in Table 1.
structure, which, according to the numerical calculations
of Augereau & Papaloizou (2004), can be produced by the
gravitative perturbation of HD 141569B and C. The dust gap 3. Forces
at 250 AU may be caused by a planet of approximately Jovian
In circumstellar nebulae, solid bodies experience a number of
mass, but it is not yet clear whether gas giants can form within
perturbing forces in addition to gravity, causing them to move
a few million years at a distance of several hundred AU from
along non-Keplerian orbits. In a first approximation, we can as-
the star (Wyatt 2005).
sume that they describe circular orbits, the radii of which shrink
or grow in the course of time, depending on the size of the body
Both disks – HR 4796A and HD 141569A – surround lumi- and its distance from the star. We now introduce the different
nous, early-type stars. Since we wish to explore the dependence forces, presenting an analytic expression for each.
of the particle dynamics on the stellar luminosity as well, we
add two more, hypothetical systems with gas densities equal
3.1. Photophoresis
to those of HR 4796A and HD 141569A, but around a star of
solar luminosity. They can be viewed as representations of the While Rohatschek (1996) found a semi-empirical formula for
solar nebula at different stages of dissipation (Hollenbach et al. the photophoretic force at all gas pressures, Beresnev et al.
1994; Hollenbach et al. 2000). (1993) derived an analytic expression for spherical, nonrotating
4 Herrmann & Krivov: Photophoresis in transitional disks

objects with a homogenous surface using an elaborate theoret- hLlG (HR 4796A)
ical approach starting from the molecular velocity distribution 1e+12
function (compare also Tehranian et al. 2001). The acceleration
1e+10
due to the photophoretic force they found is:
1e+08
π µ g mH
r
I J1
aphot =
4 s ρbulk 2kT 1e+06

Kn
αE Ψ1 10000 1 mm
× . (3)
αE + 15ΛKn (1 − αE)/4 + αE ΛΨ2
100 1m

Here, s is the particle radius, I the radiation intensity, J1 the 1


asymmetry parameter that describes the accommodation of
gas molecules to the particle’s surface and light absorption. 0.01
0.1 1 10 100
Assuming complete absorption and an accommodation prob- r / AU
ability of 100 %, we set J1 = 0.5. The mean molecular weight hLhG (HD 141569A)
of a gas of solar composition is denoted by µg = 2.34 and the 1e+08
mass of a hydrogen atom by mH . Further, k is Boltzmann’s con- 1e+07
stant, T the gas temperature and ρbulk the density of the solid 1e+06
material. For icy aggregates, ρbulk ≈ 1000 kgm−3 is a good ap- 100000
proximation. The energy accommodation coefficient αE is the 10000
fraction of molecules in contact with the surface that accommo- 1000

Kn
1 mm
date to the local temperature, which enables them to contribute 100
to photophoresis. In our work we assume complete accommo- 10 1m

dation, and thus set αE = 1. The heat exchange parameter Λ, 1

which describes the particle’s thermal relaxation properties, is 0.1

determined as 0.01
0.001
kth + 4εσ T 3 s 0.1 1 10 100
Λ= , (4) r / AU
kgas
Fig. 1. Knudsen numbers Kn = L/s for systems with low (top)
where kth is the material’s thermal conductivity, ε its emissiv- and high (bottom) gas content as a function of distance from
ity, σ the Stefan-Boltzmann constant and kgas the thermal con- the star. In each panel, four lines correspond to the particle radii
ductivity of the gas. Assuming black bodies, we set ε = 1. The of 1 mm, 1 cm, 10cm, and 1 m. Since L only depends on gas
quantities Ψ1 and Ψ2 are given by density, Kn does not vary with stellar luminosity.
!
Kn 2π 1/2 Kn
Ψ1 = 1+ ,
Kn + (5π /18) 5 Kn2 + π 1/2 Kn + π /4
! Here, p = ngas kT is the gas pressure (assuming ideal gas). In
1.21 π 1/2 Kn
 
1 15 transitional disks, we can usually assume to be in the high
Ψ2 = + Kn 1− , (5)
2 4 100 Kn2 + π /4 Knudsen number regime, except for large objects in the inner-
most parts of the disks (Fig. 1). The simplified expression (6)
and Kn ispthe Knudsen number. It is defined as Kn ≡ L/s, where will allow us to find useful approximate solutions for the parti-
L = 1/( (32) π ngas rg2 ) is the molecule’s mean free path, with cles’ stability radii (i.e. the radii of stable circular orbits). For
rg ≈ 10−10 m being the radius of gas molecules and ngas their large particles (s ≥ 10 cm) and small distances from the star
number density. (r ≤ 1 AU), the more general expression (3) will be used in nu-
Expression Eq. (3) for the photophoretic acceleration is merical calculations.
valid for all Knudsen numbers. In the case of high Knudsen The terms ϒhc , ϒrad and ϒgas are related to three different
numbers (mean free paths of molecules are large compared to processes which reduce the photophoretic effect. The first term,
particle sizes), Eq. (3) reduces to (see Krauss & Wurm (2005), ϒhc , corresponds to the transport of thermal energy through the
their Eq. 1, and Beresnev et al. (1993), their Eq. 26) object – i.e. heat conductivity, which reduces the temperature
gradient, thus lowering the efficiency of photophoresis. The
I pJ1 second one, ϒrad , describes thermal radiation from the parti-
aphot = , (6)
4ρbulk (ϒhc + ϒrad + ϒgas ) cles’ surface. Since it is proportional to T 4 , much more energy
where is radiated from the warm than from the dark side, which re-
sults in a substantial reduction of the temperature gradient, and,
ϒhc = kth T , (7) therefore, of photophoresis. Obviously, this process gains effi-
ϒrad = 4σ T 4 ε s , (8) ciency with increasing temperature. Finally, the third term ϒgas
q describes heat conduction away from the object’s surface into
ϒgas = p 2kT /π µg mH s . (9) the surrounding gas. Since it also shows a weak dependence
Herrmann & Krivov: Photophoresis in transitional disks 5

on temperature (∝ T 1/2 ), it causes a reduction of temperature with Qpr being the radiation pressure efficiency. We will make
gradient too. use of the constant B later. The effective ‘photogravitational’
Krauss & Wurm (2005) and Wurm & Krauss (2006) use acceleration is then
only the first term ϒhc in the denominator of Eq. (6), which is a Meff
good approximation for particles with s ≪ 1 mm and only in the agrav, eff = −G r. (12)
r3
case of low luminosity (i.e. solar-type) systems (see Sect. 5.2).
In order to analyze the relative importance of the three ‘coun- In our work, we simply set Qpr = 1, assuming the solids to
tereffects’ ϒhc , ϒrad and ϒgas , we plot them for HD 141569A as be black bodies. The reason is that generalization to non-black
functions of particle size (Fig. 2) at three different distances. surfaces requires a change not only in the emissivity, but also in
the asymmetry parameter J1 . The latter is not known for real-
istic materials, and its determination requires complicated nu-
1000 merical procedures (see, e.g. Mackowski 1989). Analytical
100 solutions are possible, but only in the cases of high or low ab-
10
sorption coefficients (see Arnold & Lewittes 1982).
1

0.1 3.3. Gas drag force


[W m-1]

0.01 Unlike the other three forces (gravity, radiation pressure, pho-
0.001 tophoresis), gas drag is not a radial force. Since it results from
1e-04 the momentum transferred to the body by molecules impinging
1e-05 on it as it moves through the gas, the force vector is antiparallel
Υhc to the relative velocity ∆v of the dust grain with respect to the
1e-06 Υrad
Υgas gas.
1e-07
100 1000 10000 100000 1e+06 To calculate gas drag, we first need a model for the motion
s / µm of the gas component of the circumstellar nebula. Since the
pressure gradient supports gas against stellar gravity, it travels
Fig. 2. Quantitative comparison of the three reduction pro-
on circular orbits with a sub-Keplerian speed
cesses for hLhG (HD 141569A) at r = 5, 10, 20 AU (from
vg = vK 1 − η ,
p
above for each set of lines). For the other systems the result (13)
looks qualitatively similar, with ϒgas being in the case of the
gas-poorer systems completely negligible for all sizes s. and angular velocity
Ωg = ΩK 1 − η ,
p
(14)
Figure 2 shows that for fairly small particles (s ≤ 1 mm),
where vKp= GM⋆ /r is the Keplerian circular velocity, ΩK =
p
heat conduction (ϒhc ) is the main reduction process. At larger
sizes (s ≥ 10 cm), radiation (ϒrad ) takes over. In the intermedi- vK /r = GM⋆ /r3 the corresponding angular velocity, and η
ate interval, both processes contribute to the reduction of pho- is the ratio of pressure gradient force to gravity:
tophoresis. At yet larger sizes (s ≥ 1 m), heat conduction from 1 dp
the surface into the surrounding gas (ϒgas ) comes into play, ap- η =− . (15)
rΩ2K ρg dr
proaching values similar to that of ϒhc . As lower gas pressure
dramatically reduces the importance of ϒgas , it is completely The η ratio can be rewritten as:
negligible for all object radii in the case of the gas-poor sys-
µg −1 L⋆ 1/4
   
1
tems hLlG (HR 4796A) and lLlG. η = 1.1 × 10−3 −q
2 2.34 L⊙
 −1 
M⋆ r  1/2  r 1/2
3.2. Radiation pressure × ≡E . (16)
M⊙ 1 AU 1 AU
Since radiation pressure is proportional to r−2 ,
it can be
If the motion of the particles is subsonic (∆v ≪ vT ≈ cS
taken into account by introducing an ‘effective stellar mass’
where cS is the speed of sound – this is always the case in the
(Burns et al. 1979):
transitional disks), the gas drag acceleration is given by (see
Meff = M⋆ (1 − β ) , (10) Takeuchi & Artymowicz 2001):
3ρg
where β stands for the ratio of accelerations due to radiation aD = − vT ∆v , (17)
4ρbulks
pressure and gravity:
where
arad 0.5738 Qpr L⋆ /L⊙
β= = 4

8kT
1
2 4
agrav ρbulk[g cm−3 ] s[µ m] M⋆ /M⊙ vT = = × hvtherm i (18)
 −1 3 π µ g mH 3
s
≡B , (11)
1 µm is 4/3 times the mean thermal velocity.
6 Herrmann & Krivov: Photophoresis in transitional disks

The reaction of particles to the gas drag force critically de- In the first case, the radial drift velocity vr ≡ dr/dt computes to
pends on their size. Small objects adjust their velocity instan- (see Weidenschilling 1977, sect. 4.1, 4.2):
taneously, they are swept along with the gas component. Large
vr, small = ts ∆g = Ts (β + χ − η ) rΩK . (25)
ones react only sluggishly to gas drag, taking a longer time to
change their velocity substantially. In order to create a quan- In the second case, the orbit decays at a rate
titative measure for the tendency of solids to be influenced by r 2 r ∆g (β + χ − η ) rΩK
gas drag, the stopping time is introduced. We denote the time vr, large = − ∆v = = . (26)
ts vK ts agrav Ts
needed for a particle injected into the gas to be slowed down to
e−1 times its initial velocity by ts = ∆v / aD . The dimensionless These two formulas can be combined into (see
stopping parameter Ts is defined by: Takeuchi & Artymowicz 2001, Sect. 3.3):
4ρbulksvK β +χ −η
Ts ≡ ts ΩK ≈ . (19) vr = vK . (27)
3ρg rvT Ts + Ts−1
The right-hand side approximation holds in the case of sub- In Fig. 3, we plot the absolute value of vr over s for r =
sonic motion. It renders the stopping time independent of the 10 AU, both with and without photophoresis. The points where
particle’s momentary velocity. vr = 0 are shifted to larger sizes by photophoresis. Note that for
larger sizes, radial velocities depend almost exclusively on gas
density since for them gas drag (η in Eq. 27) becomes the main
4. Equation of motion driving force of radial migration.
The equation of motion of a dust particle is:
d2 r No Photophoresis
= agrav, eff + aphot + aD . (20) 1000
dt 2 hLlG (HR 4796A)
hLhG (HD 141569A)
100
Being interested in the radial motion of the bodies, we now lLlG (later SN)
lLhG (earlier SN)
consider the radial component of the equation of motion (20). 10
In the reference frame corotating with the gas, it takes the form 1
|vr| / m s-1

d2 r 0.1
− Ω2g r = −agrav, eff + aphot . (21)
dt 2 0.01

The acceleration ≡ ∆g of the particle in the corotating


d2 r/dt 2 0.001
frame can be interpreted as the ‘residual gravity’. It computes
1e-04
to
1e-05
∆g = −agrav, eff + aphot + Ω2g r = (β + χ − η ) Ω2K r , (22) 100 1000 10000 100000 1e+06 1e+07 1e+08
s / µm
where χ ≡ aphot /agrav is the photophoresis-to-gravity ratio. Photophoresis
Particles experience the inward-directed residual acceleration 1000
hLlG (HR 4796A)
η Ω2K r and the outward-directed accelerations due to pho- 100
hLhG (HD 141569A)
lLlG (later SN)
tophoresis (χ Ω2K r) and radiation pressure (β Ω2K r). lLhG (earlier SN)
10
Assuming now that the solid particle moves at a Keplerian
circular speed, we can write 1
|vr| / m s-1

0.1
v2g v2
= K + ∆g . (23) 0.01
r r
The relative velocity is then approximately 0.001

∆g 1e-04
 
∆v = vK − vg ≈ − vK . (24) 1e-05
2 agrav 100 1000 10000 100000 1e+06 1e+07 1e+08
s / µm
We now use ∆g to derive an approximation for the ra-
dial drift velocity of the particles. We can discern two limiting Fig. 3. The absolute values of the radial velocities at r = 10 AU.
cases: The minima correspond to the sizes for which vr = 0.
– Small particles are swept along with the gas. They stay on
circular orbits, while moving with sub-Keplerian velocity, In what follows, differential equation (27) for r(t) will be
which causes them to experience the residual gravity pull studied analytically and solved numerically. In the realm of
∆g. large radial velocities (or equivalently, small particles), r(t)
– Large particles are decoupled from gas. They move on should be computed from the ‘exact’ equation of motion (20).
Keplerian orbits, experiencing a head wind that gradually However, our numerical tests have shown Eq. (27) to be accu-
reduces their angular momentum. rate enough for all particle sizes larger than ∼ 100 µ m.
Herrmann & Krivov: Photophoresis in transitional disks 7

5. Results 5.2. Equilibrium distance


We now calculate the equilibrium distance which, according to
5.1. Radial motion Eq. (27), must satisfy
β + χ − η = 0. (28)
We used Eq. (27) to calculate the radial drift of different-sized
particles. What happens, in short, is that the particles migrate From this, an implicit expression for the radius rstab (s) of the
inward or outward, until they reach stable circular orbits on stable orbit can be derived:
which gravity, centripetal force (in the particle’s own inertial
v2K (rstab ) (1 − η (rstab ))
system), photophoresis and radiation pressure balance each rstab (s) = . (29)
other, thus permitting circular motion (vr = 0), while at the agrav, eff (rstab , s) − aphot(rstab , s)
same time ∆v = 0, i.e. the particle travels at the same speed
as the gas.
High luminosity

Phot. (low gas content)


No Phot. (low gas)
Phot. (high gas)
No Phot. (high gas)
Photophoresis 10
No Photophoresis
s = 500 µm

rstab / AU
10

1
r / AU

s = 1 cm s = 3 cm

1 s = 3 mm

100 1000 10000 100000 1e+06 1e+07


s / µm
Low luminosity

0 2000 4000 6000 8000 10000 12000 Phot. (low gas)


t / yr Phot. (high gas)
10 No Phot. (high + low gas)

Fig. 4. Distance as a function of time for particles of four dif-


ferent radii in the hLlG system (HR 4796A). Dashed lines:
rstab / AU

without photophoresis, solid: with photophoresis. The lower


1
border of the plot corresponds to the evaporation limit revap =
0.16 AU (defined by black-body equilibrium temperature T =
1500 K).
0.1

10 100 1000 10000 100000 1e+06 1e+07


As an example, Fig. 4 shows the r(t)-curves for particles s / µm
of four different sizes with and without photophoresis in the
system hLlG (HR 4796A). The particles start on circular or- Fig. 5. Stability radii as functions of particle radius s for two
bits with r0 = 5 AU. In fact, initial circularity is not important, high-luminosity systems (top) and two low-luminosity ones
because elliptic orbits are quickly circularized by the gas drag (bottom). In the bottom panel, the ‘No Photophoresis’ curve
force. The initial value r0 does not matter either, because the is identical for both systems, as they differ only in the value
particles migrate towards their equilibrium orbits within several of gas density, which does not affect the stability condition if
thousand years. While small, outward migrating bodies travel photophoresis is not taken into account (η only depends on the
slowly towards their equilibrium distance rstab (s), approach- density exponent q, not its absolute value – see Eq. 16). Again,
ing it asymptotically, the larger, inward migrating ones almost the lower borders of the plots correspond to the evaporation
‘drop’ onto their stability orbits, being stopped almost instanta- limits.
neously. Note, however, that ‘instantaneous’ here refers to a de-
celeration process taking several centuries. Therefore, the large The equilibrium distance rstab (s) depends on the system’s
particles travel essentially on Keplerian orbits, corresponding parameters as well as on the particle size. Fig. 5 shows solu-
to the second case described in sect. 4. With increasing particle tions of Eq. (29), computed with and without photophoresis. It
size, though, this process takes longer, as larger objects experi- can be clearly seen that photophoresis significantly increases
ence weaker drag acceleration than smaller ones. Another con- the radii of stable orbits – as expected: photophoresis pushes
clusion from Fig. 4 is that the stability radii are pushed outward particles away from the star. The curves exhibit a character-
by photophoresis. istic shape: after branching off from the solutions computed
8 Herrmann & Krivov: Photophoresis in transitional disks

without photophoresis, they flatten. While for hLhG there is change M⋆ accordingly. Analyzing the plots in Fig. 7, we can
only a small region in which the curve is flatter, the two low- identify the following dependencies:
luminosity systems have well-developed plateaus, the length
and distance from the star of which increase with gas den- – Higher gas density ρ0 moves the curves’ middle parts to
sity. The reason for this is explained below. Note also, that larger r, which is, of course, a consequence of photophore-
the plateaus in the rstab (s)-curve correspond to areas, where sis being directly dependent on gas pressure. Also, the lim-
solids have a higher surface density. At a certain particle size, iting particle size beyond which no stable orbits exist de-
the curves drop to zero suddenly. This is a consequence of pends almost exclusively on ρ0 , while the influence of the
mean free path becoming smaller than particle sizes – in this other parameters is rather weak.
area, Eq. (6) is not a good approximation anymore, and Eq. (3) – The density exponent q, which determines the slope of ρ (r),
should be used instead to calculate the photophoretic accelera- also controls the slope of rstab (s): a lower absolute value
tion. Because in the low Knudsen number regime photophore- of q produces a steeper curve, as the gas disk spreads to a
sis decreases in strength as body size increases, Eq. (28) and greater distance from the star.
(29) do not have solutions above a certain body radius, which – Stellar luminosity L⋆ determines the characteristic shape
causes rstab (s) to drop to zero, so that there are no stable orbits of the rstab (s) curve. While large luminosities produce a
for solids above this critical size. These considerations are visu- monotonically decreasing curve, lower (i.e. solar) ones gen-
alized in Fig. 6, where η and χ are plotted as functions of r for erate a plateau at a certain distance from the star, which
different radii s. The third force parameter, β , can be neglected gives rise to a concentration of particles of different sizes
for large particles (s ≫ 100 µ m), and therefore is not shown. in that region.
We can identify two cases. For small bodies, the χ (r)-curve The mechanism behind the third effect is to be found in the
has two intersections with η (r). This means that in a certain denominator of Eq. (6). As we have seen in Fig. 2, in different
size range, (28)–(29) have got two solutions. In our work we size ranges three different processes reduce the photophoretic
consider only the one farther out from the star, since in transi- effect – heat conduction in small particles, radiation for larger
tional disks we expect most particles to drift inward from larger ones and heat conduction into the surrounding gas for very
radii, where they are produced by collisions between left-over large objects. If the stellar luminosity is low, radiation becomes
planetesimals in exo-Kuiper belts. Besides, for three out of four important only for relatively large sizes, while for smaller ob-
systems (hGlL is the exception), the inner solution lies in the jects only ϒhc is relevant in Eq. (6). As ϒhc is independent of
sublimation zone. As s increases further, the χ (r)-curve moves size, the resulting curve tends to run parallel to the s-axis, cre-
downward until it no longer crosses the η -curve, and thus no ating the plateau.
solution of (28)–(29) remains. This means that above a certain In the following, we formulate these considerations quan-
size no stable orbits exist. titatively. We are going to compute solutions of Eq. (29)
for very large and very small particles, using the simplified
large-Knudsen approximation for the photophoretic accelera-
1 tion given in Eq. (6). We need to calculate the following ratios:
η
χ (s = 1 mm)
χ (s = 3 cm) aphot, ϒhc
0.1 χ (s = 30 cm) χ (ϒhc ) = (30)
evaporation limit agrav
0.01
and
aphot, ϒrad
0.001 χ (ϒrad ) = , (31)
agrav
1e-04
where aphot, ϒhc and aphot, ϒrad denote the accelerations due to
(large Knudsen regime) photophoretic force using exclusively
1e-05
ϒhc and ϒrad in the denominator, respectively. The first case is
a good approximation for small particle radii (s ≤ 1 mm, com-
1e-06
0.01 0.1 1 10 pare Fig. 2), the second one for larger, but sub-meter objects
r / AU (1 mm ≤ s ≤ 1 m).
Fig. 6. The force ratios η and χ as functions of r for different The quantities χ (ϒhc ) and χ (ϒrad ) can be computed from
the following expressions:
particle sizes, in the hLlG system (HR 4796A). The number of
intersections between the curves corresponds to the number of 
M⋆ −1 L⋆  µg −1
  
solutions of (28)–(29). χ (ϒhc ) = 1.0 × 10 8
M⊙ L⊙ 2.34
−1 
ρd

r q
× −3
ρ0
1000 kgm 1 AU
Next, we determine how the exact shape of rstab (s) is influ-  r q
enced by the choice of system parameters. Therefore, in Fig. 7, ≡ Z(ϒhc ) (32)
we plot the stability radii for different choices of L⋆ , ρ0 and q. 1 AU
Note that, when varying L⋆ in the bottom panel of Fig. 7, we and
Herrmann & Krivov: Photophoresis in transitional disks 9

ρ0 = 3.12 ... 156 * 10-10 kg m-3, q = -2.25, L = 22.4 Lsun, M = 2.3 Msun It is easy to show that in the first (small-particle) case, Eq.
rstab
(29) can be solved for s:

high ρ0
sstab, small (r) B
= . (34)
1 µm 1/2
E (r [AU]) − Z(ϒhc ) (r [AU])q
10 low ρ0
For definition of B and E, see Eqs. (11) and (16). This allows
rstab / AU

us to calculate the size of particles on stable orbits at a certain


stellar distance r. Since q is negative, sstab, small tends to infinity
as r approaches zero. This produces the curve’s plateau, i.e. the
1 concentration belt. From Eq. (34), the belt’s radius computes
to:
 1/(q−1/2)
100 1000 10000 100000 1e+06
E
rbelt = (35)
s / µm Z(ϒhc )
ρ0 = 156 * 10-10 kg m-3, q = -1.5 ... -3.0, L = 22.4 Lsun, M = 2.3 Msun
For the four systems considered, the resulting distances are
rstab listed in Table 2.
small |q|

10
Table 2. Existence and parameters of the belt.
large |q|
rstab / AU

System rbelt /AU C


lLlG (‘older’ SN) 2.0 0.8
lLhG (‘younger’ SN) 6.5 2.6
hLlG (HR 4696A) 4.0 0.1
1
hLhG (HD 141569A) 22.8 0.07

100 1000 10000 100000 1e+06 In the second (large-particle) case, the corresponding ex-
s / µm pression
ρ0 = 156 * 10-10 kg m-3, q = -2.25, L = 1 ... 21 Lsun, M = 1.0 ... 2.3 Msun
sstab, large (r) B (r [AU])−1/2 + Z(ϒrad ) (r [AU])q+1
rstab = (36)
1 µm E
can be derived.
high L The condition for the formation of the belt is that at its
rstab / AU

position, the value of sstab, large has to be sufficiently larger


10
than sstab, small . If that is the case, the asymptotic behavior of
sstab, small is seen in the ‘exact’ curve too, otherwise it is over-
low L
ridden by sstab, large (see Fig. 8). To evaluate this condition nu-
merically, we set

rbelt = rbelt + ∆r (37)
10 100 1000 10000 100000
s / µm where for ∆r a sufficiently small value has to be chosen. We
use ∆r = 0.1 AU. This is necessary because sstab, small cannot be
Fig. 7. Dependence of the equilibrium distance rstab (s) on the
computed at rbelt , as the denominator of Eq. (34) becomes zero
gas density ρ0 , the density exponent q, and stellar luminosity
at this point. Then, the ratio of the two approximating functions
L⋆ . Top: ρ0 running, q and L⋆ fixed. Middle: q running, ρ0 and ′
at rbelt can be used as a measure of how pronounced the belt is:
L⋆ fixed. Bottom: L⋆ running, q and ρ0 fixed. The parameter
values are indicated on top of each panel. ′ ′
C ≡ sstab, large (rbelt ) / sstab, small (rbelt ). (38)
The values of C are listed in Table 2. Comparing with Fig. 5, we
find that for C ≪ 1, the plateau does not appear at all, or is only
marginal. For C ≈ 1, the belt is well developed. The particle
M⋆ −1 L⋆ 1/4  µg −1
  

size range it encompasses and its degree of concentration (i.e.
χ (ϒrad ) = 2.1 × 1010 the ’flatness’ of the curve in that area) increase with C, which
M⊙ L⊙ 2.34
−1  −1 makes it a direct measure for the system’s tendency to produce
ρd
 
r q+3/2 s
× ρ0 a belt. In fact, other choices for ∆r are possible – for them, the
1000 kgm−3 1 AU 1 µm values with which C has to be compared, have to be changed.
 r q+3/2  s −1 For ∆r = 0.1 AU, the critical value above which well-formed
≡ Z(ϒrad ) . (33)
1 AU 1 µm belts appear is C = 1.
10 Herrmann & Krivov: Photophoresis in transitional disks

100
hLlG (HR 4796A)
hLhG (HD 141569A)
lLlG (older SN)
lLhG (younger SN)
10
10
rstab / AU

rstab(s) (lLhG)
rstab small(s) (lLhG) 0.1
1 rstab large(s) (lLhG)
rstab(s) (hLhG)
rstab small(s) (hLhG)
rstab large(s) (hLhG)
0.01
10 100 1000 10000 100000 1e+06 1 10 100 1000 10000 100000 1e+06
s / µm s / µm

Fig. 8. ‘Exact’ function rstab (solid lines) and its two approx- Fig. 9. Ratio of radial velocities, computed with and without
imations (dashed lines: Eq. 34; dotted: Eq. 36), for the lLhG photophoresis. The stronger the deviation from unity, the more
(thick lines) and the hLhG (thin) systems. Note how the be- important is photophoresis. The stellar distance is r = 10 AU
havior of rstab, small produces the curve’s plateau – in the higher
luminosity case, rstab, large ‘takes over’ at smaller particle sizes, Table 3. Size range of particles affected by photophoresis.
thus keeping rstab, small from producing the effect.
System smin smax
lLlG (‘older’ SN) 9 µm 3 cm
lLhG (‘younger’ SN) 1 µm 1.5 m
We see that low luminosities and high gas densities are cru- hLlG (HR 4696A) 10 µ m 3.2 cm
cial for belt formation. Since ϒrad ∝ T 4 ∝ L⋆ , with rising lumi- hLhG (HD 141569A) 1 µm 6m
nosity ϒrad becomes important at smaller particle sizes, super-
seding ϒhc before the plateau of sstab, small is reached. Because
they increase the overall strength of photophoresis, higher gas
densities push the plateau outward, reducing its slope – and 6. Rotation of particles
thus increasing the degree of particle concentration in the belt. Until now, we considered only nonrotating particles. We have
Fig. 8 demonstrates how the exact solution of Eq. (29) is to check whether this assumption is realistic, since the rotation
approximated by Eq. (34) and Eq. (36) in different particle size of bodies can transport thermal energy from the dark to the lit
ranges for the lLhG and hLhG models. side, thus banishing the photophoretic effect. There is a number
of different mechanisms that can induce particle rotation:
5.3. Particle size range
– Collisionally induced rotation. During the formation of
As we have seen, particles of different sizes are influenced bodies as well as during their life, they collide with other
by photophoresis to a different extent. Its relative importance particles, which can change their orbits as well as transfer
depends on their particle radius, declining for very small angular momentum to them, spinning them up.
(micrometer-range) and very large (above meter-range) ob- – Rotation induced by gas drag or radiation forces. If par-
jects. In between, photophoresis plays an important role. In ticles are not exactly spherical, any force acting on their
order to analyze the critical size range, we plot the ratio x ≡ surface will change their angular momentum. For instance,
|vr, phot /vr, no phot | as a function of s in Fig. 9, where vr, phot and radiation pressure may spin-up the particles (the so-called
vr, no phot are the radial velocity with and without photophoresis windmill effect, Paddack & Rhee 1975), but may also sta-
respectively. The distance is set to r = 10 AU. Note that the in- bilize/align then (Draine & Weingartner 1996). Such ef-
terval between the two peaks corresponds to the size range in fects are beyond the scope of this paper.
which the two vr have got different signs, because the stability
radius with photophoresis is greater than 10 AU, while the one Whether rotation is able to subdue photophoresis depends on
calculated without photophoresis is smaller. four different timescales: tcoll , the typical time between two col-
If we use the criterion that |x − 1| ≥ 0.001, photophore- lisions; ts , the gas coupling time which also determines the time
sis has to be taken into account in the size ranges listed in needed to slow down rotation; theat , the thermal relaxation time
Table 3. Note, however, that for s ≤ 10 µ m, the above evalu- needed to establish a stable heat gradient within the particle;
ation is no longer necessarily valid, as for small particles, the and trot , the typical rotation period.
photophoretic force can reverse its direction (Tehranian et al. The collisional time tcoll can be estimated in a standard way:
2001). Also, for hLhG (HD 141569A), β (smin ) is greater than
0.5, thus these particles are β -meteoroids which escape from 1
tcoll = , (39)
the system. np vσcoll
Herrmann & Krivov: Photophoresis in transitional disks 11

where np is the particle number density, v the collisional veloc- 1e+06


High luminosity
ity and σcoll = π (s1 + s2 )2 is the collisional cross section for Low luminosity
spherical particles of radii s1 and s2 . With
ρd
100000
np = , (40)
ρbulk 43 π s3
10000
and the standard assumption ρd /ρg = 10−2, we get for particles
of equal size (cf. Krauss et al. 2006, their Eq. 15):
ρbulks 1000
tcoll = 33 . (41)
ρg v
The stopping time ts was computed in section 3.3, Eq. (19). 100
0.1 1 10 100
Finally, the thermal relaxation time theat is given by (see
v / m s-1
Krauss & Wurm 2005, Eq. 6):
1e-04
hLlG (HR 4796A)
ρbulkcd s2 hLhG (HD 141569A)
lLlG (later SN)
theat = , (42) lLhG (earlier SN)
kth 1e-05

where cd ≈ 1000 Wm−1K−1 is the particle’s heat capacity.


1e-06
We can use two different approaches to deal with rota-
tion. One is to compare collision time to stopping time: if
tcoll /ts ≫ 1, damping of rotation occurs faster than excitation by 1e-07
impacts, and therefore photophoresis is not significantly weak-
ened. Using Eqs. (41) and (19), we get 1e-08

tcoll vth
= 25 . (43)
ts v 1e-09
100 1000 10000 100000 1e+06 1e+07 1e+08
Another option is to check whether rotation, once it has been s / µm
excited by collisions, is sufficiently slow to allow a stable heat
Fig. 10. Ratios of typical timescales. Top: tcoll /ts as a function
gradient to be established within the particle: this is the case
of v. Bottom: trot /theat as a function of s.
if trot /theat ≫ 1. In order to find a rough estimate for trot , we
assume that the particle transforms the entire kinetic energy of
its radial motion vr into rotational energy: end, we employed a single-body approach and added the pho-
tophoretic force to the standard array of perturbing forces (stel-
1 2π 2
 
1
Jsphere = m v2r , (44) lar gravity, direct radiation pressure, and gas drag). Particle-
2 trot 2 particle interactions as well as particle growth were not in-
with the solid sphere’s moment of inertia Jsphere = (2/5)ms2 , cluded. We find that photophoresis may lead to noticeable cor-
we get rections to the results obtained with models that do not take it
into account.
8 πs
r
Our main results can be summarized as follows:
trot = . (45)
5 vr
1. Both with and without photophoresis, solid objects migrate
Fig. 10 shows tcoll /ts as a function of v and trot /theat as inward or outward, until they reach the stability distance
a function of s. The distance is chosen to be r = 10 AU. rstab (s), where s is the particle radius. At that distance, ra-
We see that collisional timescales are longer than stopping dial forces cancel each other in the particle’s own inertial
timescales by many decades, thus allowing rotation to decline system, and the orbital velocity is equal to that of the gas.
between collisions. Nonetheless, rotation frequencies can be The stability distance is a decreasing function of s, there-
high enough to cancel out stable temperature gradients. We fore particles are sorted according to size, with larger bod-
conclude that photophoresis can be suppressed directly after ies accumulating closer to the star. These results fully agree
collisions through rapid rotation, but remains effective for most with those by Takeuchi & Artymowicz (2001).
of the time because gas drag damps spin faster than collisions 2. Photophoresis increases the stability radii, moving objects
can excite it. to larger radial distances. The effect is noticeable in the
size range from several micrometers to several centime-
ters (for older transitional disks) or even several meters (for
7. Conclusions and discussion
younger, more gaseous, ones).
In this paper, we have analyzed the effect of photophoresis on 3. The steady-state distribution of solids is completely charac-
the dynamics of solid particles in the optically-thin, yet suffi- terized by the function rstab (s), the shape of which depends
ciently gas-rich, transitional disks around young stars. To this on the system’s parameters:
12 Herrmann & Krivov: Photophoresis in transitional disks

– Higher gas densities move the curve to larger distances Acknowledgments. We wish to thank Gerhard Wurm for
without changing its overall shape. Gas density also useful discussions, Jürgen Blum and Bo Gustafson for pointing
controls the maximum particle size up to which stable out several papers on the photophoretic effect and the anony-
orbits can exist. mous referee for helpful comments. F.H. is supported by the
– The steepness of the gas density radial profile deter- graduate student fellowship of the Thuringia State.
mines the slope of rstab (s), with flatter profiles gener-
ating steeper curves.
References
– Stellar luminosity determines the curve’s shape. While
high luminosities (∼ 20 L⊙ ) produce simple decreasing Alibert, Y., Mordasini, C., Benz, W., & Winisdoerffer, C. 2005,
curves, low (solar) luminosities generate a plateau at a A&A, 434, 343
distance from the star that can be computed with the aid Ardila, D. R., Lubow, S. H., Golimowski, D. A., et al. 2005,
of Eq. (35). In this area, objects in a certain size range ApJ, 627, 986
may accumulate. Arnold, S. & Lewittes, M. 1982, J. of Appl. Phys., 53, 5314
4. Particle rotation tends to reduce the photophoretic effect. Augereau, J.-C. 2006, in: Semaine de l’Astrophysique
Our estimates suggest, however, that it is damped by gas Francaise, ed. D. Barret, F. Casoli, G. Lagache,
drag quickly enough to keep photophoresis at work. A. Lecavelier, & L. Pagani, 375
Augereau, J. C., Lagrange, A. M., Mouillet, D., & Ménard, F.
Our work predicts the formation of a particle concentration 1999a, A&A, 350, L51
belt at a certain distance from the star. For the high luminos- Augereau, J.-C., Lagrange, A.-M., Mouillet, D., Papaloizou, J.
ity systems, it is not very pronounced (hLhG) or does not ap- C. B., & Grorod, P. A. 1999b, A&A, 348, 557
pear at all (hLlG). Furthermore, the radii of the observed rings Augereau, J. C. & Papaloizou, J. C. B. 2004, A&A, 414, 1153
around HD 141569A are an order of magnitude larger than that Balog, Z., Muzerolle, J., Rieke, G. H., et al. 2007, ApJ, 660,
of the slight concentration belt predicted for the hLhG system. 1532
It is not likely therefore that the observed structures around Beresnev, S., Chernyak, V., & Fomyagin, G. 1993, Phys. Fluids
HR 4796A and HD 141569A are caused by photophoresis. A, 5, 2043
While photophoresis is probably active in transitional disks, Besla, G. & Wu, Y. 2007, ApJ, 665, 528
the circumstellar rings of HR 4796A and HD 141569 must Blum, J. & Wurm, G. 2000, Icarus, 143, 138
be shaped by other forces and effects, such as gravitational Bouwman, J., Meyer, M. R., Kim, J. S., et al. 2006, in
sculpting by planets or interactions with stellar companions. Planet formation: Theory, observations and experiments, ed.
Alternatively, a rapid decline of gas density at the disks’ edges H. Klahr & W. Brandner, 14
(Takeuchi & Artymowicz 2001) or a recently proposed dust- Brauer, F., Dullemond, C. P., Johansen, A., et al. 2007, A&A,
gas instability (Besla & Wu 2007) may cause particles to accu- 704, 1169
mulate there. Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1
The model presented here is rather exploratory and rests on Calvet, N., D’Alessio, P., Watson, D. M., et al. 2005, ApJ, 630,
a number of simplifying assumptions. In the future, we plan to L185
investigate the problem a more realistic way, lifting some of Chen, C. H. & Kamp, I. 2004, ApJ, 602, 985
the assumptions we made to keep the problem tractable. First, Currie, T., Balog, Z., Kenyon, S. J., et al. 2007, ApJ, 659, 599
we plan to deal with particle-particle interactions, taking col- Draine, B. T. & Weingartner, J. C. 1996, ApJ, 470, 551
lisions and growth into account. This can be done in the style Ehrenhaft, F. 1917, Phys. Z., 18, 352
of Krauss et al. (2006), increasing the radius through an expo- Eisner, J. A., Chiang, E. I., & Hillenbrand, L. A. 2006, ApJ,
nential ansatz s = s0 exp(t/t0 ). The latter corresponds to the 637, L133
assumption that the object moving through the nebula collects Fisher, R. S., Telesco, C. M., Piña, R. K., Knacke, R. F., &
smaller particles on its surface. A more detailed approach will Wyatt, M. C. 2000, ApJ, 532, L141
employ statistical methods (Krivov et al. 2005, 2006). Then, Haisch, Jr., K. E., Jayawardhana, R., & Alves, J. 2005, ApJ,
we wish to explore rotation in greater detail, calculating colli- 627, L57
sion timescales and rotation frequencies using models for two- Hayashi, C., Nakazawa, K., & Nakagawa, Y. 1985, in
body collisions. Also, the effects of gas drag and radiation Protostars and Planets II, ed. D. C. Black & M. S. Matthews,
forces on non-spherical objects need to be taken into account 1100
(see e.g. Xu et al. 1999). Other issues include the variation of Henning, T., Dullemond, C. P., Wolf, S., & Dominik, C. 2006,
physical parameters (density, thermal conductivity, emissivity, in Planet formation: Theory, observations and experiments,
J1 ) with size and distance from the star, as well as the global ed. H. Klahr & W. Brandner, 112
evolution and clearing of dust in the system which defines the Hollenbach, D., Gorti, U., Meyer, M., et al. 2005, ApJ, 631,
time around which photophoresis can come into play. 1180
In spite of these unknowns, we have demonstrated that pho- Hollenbach, D., Johnstone, D., Lizano, S., & Shu, F. 1994, ApJ,
tophoretic force in transitional circumstellar disks cannot be 428, 654
neglected, and has to be included in elaborate models of such Hollenbach, D. J., Yorke, H. W., & Johnstone, D. 2000, in
systems. Protostars and Planets IV, ed. V. Mannings, A. Boss, &
S. Russell (University of Arizona Press), 401
Herrmann & Krivov: Photophoresis in transitional disks 13

Hueso, R. & Guillot, T. 2005, A&A, 442, 703 Strom, S. E. & Edwards, S. 1993, in Planets Around Pulsars,
Jayawardhana, R., Coffey, J., Scholz, A., Brandeker, A., & van ed. J. A. Phillips, S. E. Thorsett, & S. R. Kulkarni (ASP
Kerkwijk, M. H. 2006, ApJ, 648, 1206 Conf. Ser., vol. 36), 235
Jayawardhana, R., Fisher, S., Hartmann, L., et al. 1998, ApJ, Strom, S. E., Edwards, S., & Skrutskie, M. F. 1993, in
503, L79 Protostars and Planets III, ed. E. H. Levy & J. I. Lunine,
Jura, M., Malkan, M., White, R., et al. 1998, ApJ, 505, 897 837
Jura, M., Zuckerman, B., Becklin, E. E., & Smith, R. C. 1993, Takeuchi, T. & Artymowicz, P. 2001, ApJ, 557, 990
ApJ, 418, L37 Takeuchi, T., Clarke, C. J., & Lin, D. N. C. 2005, ApJ, 627, 286
Kastner, J. H., Huenemoerder, D. P., Schulz, N. S., et al. Tehranian, S., Giovane, F., Blum, J., Xu, Y.-L., & Gustafson,
2001, in Young Stars Near Earth: Progress and Prospects, B. Å S. 2001, Int. J. of Heat and Mass Transfer, 44, 1649
ed. R. Jayawardhana & T. Greene (ASP Conf. Ser., vol. 244), Telesco, C. M., Fisher, R. S., Piña, R. K., et al. 2000, ApJ, 530,
159 329
Klahr, H. H. & Lin, D. N. C. 2000, in Disks, Planetesimals Thébault, P. & Augereau, J.-C. 2005, A&A, 437, 141
and Planets, ed. F. Garzón, C. Eiroa, D. de Winter, & T. J. van den Ancker, M. E., de Winter, D., & Tjin A Djie, H. R. E.
Mahoney (ASP Conf. Ser., vol. 219), 375 1998, A&A, 330, 145
Kley, W., Papaloizou, J. C. B., & Lin, D. N. C. 1993, ApJ, 416, Wahhaj, Z., Koerner, D. W., Backman, D. E., et al. 2005, ApJ,
679 618, 385
Koerner, D. W., Ressler, M. E., Werner, M. W., & Backman, Weidenschilling, S. J. 1977, MNRAS, 180, 57
D. E. 1998, ApJ, 503, L83 Weidenschilling, S. J. & Cuzzi, J. N. 1993, in Protostars and
Krauss, O. & Wurm, G. 2005, ApJ, 630, 1088 Planets III, ed. E. H. Levy & J. I. Lunine, 1031
Krauss, O., Wurm, G., Moussis, O., et al. 2006, A&A, 462, 997 Weinberger, A. J., Becklin, E. E., Schneider, G., et al. 1999,
Krivov, A. V., Löhne, T., & Sremčević, M. 2006, A&A, 455, ApJ, 525, L53
509 Weinberger, A. J., Rich, R. M., Becklin, E. E., Zuckerman, B.,
Krivov, A. V., Sremčević, M., & Spahn, F. 2005, Icarus, 174, & Matthews, K. 2000, ApJ, 544, 937
105 Wetherill, G. W. 1980, ARA&A, 18, 77
Lawson, W. A., Lyo, A.-R., & Muzerolle, J. 2004, MNRAS, Wieneke, B. & Clayton, D. D. 1983, in Chondrules and their
351, L39 Origins, ed. E. A. King, 284
Lissauer, J. J. 1993, ARA&A, 31, 129 Wurm, G. & Blum, J. 2006, in Planet formation: Theory, ob-
Lowrance, P. J., Schneider, G., Kirkpatrick, J. D., et al. 2000, servations and experiments, ed. H. Klahr & W. Brandner, 90
ApJ, 541, 390 Wurm, G. & Krauss, O. 2006, Icarus, 180, 487
Mackowski, D. W. 1989, Int J. of Heat and Mass Transfer, 32, Wyatt, M. C. 2005, A&A, 440, 937
843 Wyatt, M. C., Dermott, S. F., Telesco, C. M., et al. 1999, ApJ,
Meyer, M. R., Backman, D. E., Weinberger, A. J., & Wyatt, 527, 918
M. C. 2007, in Protostars and Planets V, ed. B. Reipurth, Xu, Y.-L., Gustafson, B. Å. S., Giovane, F., Blum, J., &
D. Jewitt, & K. Keil, 573 Tehranian, S. 1999, Phys. Rev. E, 60, 2347
Meyer, M. R. & Beckwith, S. V. W. 2000, in Lecture Notes in Zuckerman, B. & Becklin, E. E. 1993, ApJ, 406, L25
Physics, Berlin Springer Verlag, Vol. 548, ISO Survey of a Zuckerman, B. & Song, I. 2004, ARA&A, 42, 685
Dusty Universe, ed. D. Lemke, M. Stickel, & K. Wilke, 341
Morfill, G. E. 1983, Icarus, 53, 41
Morfill, G. E. 1988, Icarus, 75, 371
Mouillet, D., Lagrange, A. M., Augereau, J. C., & Ménard, F.
2001, A&A, 372, L61
Paddack, S. J. & Rhee, J. W. 1975, Geophys. Res. Lett., 2, 365
Papaloizou, J. C. B. & Terquem, C. 1999, ApJ, 521, 823
Papaloizou, J. C. B., Terquem, C., & Nelson, R. P. 1999,
in Astrophysical Discs - an EC Summer School, ed. A. J.
Sellwood & J. Goodman (ASP Conf. Ser., vol. 160), 186
Rohatschek, H. 1996, J. of Aerosol Sc., 27, 467
Safronov, V. S. 1972, Evolution of the Protoplanetary Cloud
and Formation of the Earth and Planets, NASA TTF-677
Shakura, N. I. & Sunyaev, R. A. 1973, A&A, 24, 337
Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARA&A, 25, 23
Simon, M. & Prato, L. 1995, ApJ, 450, 824
Stauffer, J. R., Hartmann, L. W., & Barrado y Navascues, D.
1995, ApJ, 454, 910
Steinbach, J., Blum, J., & Krause, M. 2004, Europ. Ph. J. E, 15,
287

You might also like